arxiv: v2 [math.dg] 30 Sep 2015

Similar documents
Kähler configurations of points

Notes 10: Consequences of Eli Cartan s theorem.

Clifford Algebras and Spin Groups

Introduction to Group Theory

LECTURE 10: THE ATIYAH-GUILLEMIN-STERNBERG CONVEXITY THEOREM

CALCULUS ON MANIFOLDS. 1. Riemannian manifolds Recall that for any smooth manifold M, dim M = n, the union T M =

Notation. For any Lie group G, we set G 0 to be the connected component of the identity.

REGULAR TRIPLETS IN COMPACT SYMMETRIC SPACES

THE ENVELOPE OF LINES MEETING A FIXED LINE AND TANGENT TO TWO SPHERES

The Symmetric Space for SL n (R)

(x 1, y 1 ) = (x 2, y 2 ) if and only if x 1 = x 2 and y 1 = y 2.

Chapter 8. Rigid transformations

Algebra I Fall 2007

SYMPLECTIC GEOMETRY: LECTURE 5

Lemma 1.3. The element [X, X] is nonzero.

The moduli space of binary quintics

A remarkable representation of the Clifford group

B 1 = {B(x, r) x = (x 1, x 2 ) H, 0 < r < x 2 }. (a) Show that B = B 1 B 2 is a basis for a topology on X.

arxiv: v1 [math.sg] 6 Nov 2015

Lecture 10: A (Brief) Introduction to Group Theory (See Chapter 3.13 in Boas, 3rd Edition)

Theorem 2. Let n 0 3 be a given integer. is rigid in the sense of Guillemin, so are all the spaces ḠR n,n, with n n 0.

(x, y) = d(x, y) = x y.

SYMPLECTIC MANIFOLDS, GEOMETRIC QUANTIZATION, AND UNITARY REPRESENTATIONS OF LIE GROUPS. 1. Introduction

Each is equal to CP 1 minus one point, which is the origin of the other: (C =) U 1 = CP 1 the line λ (1, 0) U 0

1: Lie groups Matix groups, Lie algebras

Finite affine planes in projective spaces

Basic Concepts of Group Theory

LECTURE 25-26: CARTAN S THEOREM OF MAXIMAL TORI. 1. Maximal Tori

Math Subject GRE Questions

The L 3 (4) near octagon

MAXIMALLY NON INTEGRABLE PLANE FIELDS ON THURSTON GEOMETRIES

Definition We say that a topological manifold X is C p if there is an atlas such that the transition functions are C p.

Some notes on Coxeter groups

GROUP THEORY PRIMER. New terms: so(2n), so(2n+1), symplectic algebra sp(2n)

LECTURE 5: SOME BASIC CONSTRUCTIONS IN SYMPLECTIC TOPOLOGY

Representation theory and quantum mechanics tutorial Spin and the hydrogen atom

Chapter 3. Riemannian Manifolds - I. The subject of this thesis is to extend the combinatorial curve reconstruction approach to curves

Points of Finite Order

1.1 Line Reflections and Point Reflections

How curvature shapes space

The Geometrization Theorem

Part II. Geometry and Groups. Year

Large automorphism groups of 16-dimensional planes are Lie groups

Topological properties

A LITTLE TASTE OF SYMPLECTIC GEOMETRY: THE SCHUR-HORN THEOREM CONTENTS

As always, the story begins with Riemann surfaces or just (real) surfaces. (As we have already noted, these are nearly the same thing).

Notes 2 for MAT4270 Connected components and universal covers π 0 and π 1.

1 Differentiable manifolds and smooth maps. (Solutions)

4.2. ORTHOGONALITY 161

THE EULER CHARACTERISTIC OF A LIE GROUP

ON THE CLASSIFICATION OF RANK 1 GROUPS OVER NON ARCHIMEDEAN LOCAL FIELDS

1 Hermitian symmetric spaces: examples and basic properties

The Classification of Nonsimple Algebraic Tangles

Geometric Structures in Mathematical Physics Non-existence of almost complex structures on quaternion-kähler manifolds of positive type

Stratification of 3 3 Matrices

Master Algèbre géométrie et théorie des nombres Final exam of differential geometry Lecture notes allowed

A GLIMPSE OF ALGEBRAIC K-THEORY: Eric M. Friedlander

Linear connections on Lie groups

Reduction of Homogeneous Riemannian structures

1 Differentiable manifolds and smooth maps

RIGIDITY OF MINIMAL ISOMETRIC IMMERSIONS OF SPHERES INTO SPHERES. Christine M. Escher Oregon State University. September 10, 1997

An Invitation to Geometric Quantization

The complex projective line

Möbius Transformation

Tangent spaces, normals and extrema

fy (X(g)) Y (f)x(g) gy (X(f)) Y (g)x(f)) = fx(y (g)) + gx(y (f)) fy (X(g)) gy (X(f))

Geometry of the Special Unitary Group

Particles I, Tutorial notes Sessions I-III: Roots & Weights

274 Curves on Surfaces, Lecture 4

QUALIFYING EXAMINATION Harvard University Department of Mathematics Tuesday August 31, 2010 (Day 1)

13 Spherical geometry

Part IB GEOMETRY (Lent 2016): Example Sheet 1

1 Differentiable manifolds and smooth maps. (Solutions)

Manifolds with holonomy. Sp(n)Sp(1) SC in SHGAP Simon Salamon Stony Brook, 9 Sep 2016

by Robert L. Bryant 0. Statement of the Theorem

Problems in Linear Algebra and Representation Theory

1. If 1, ω, ω 2, -----, ω 9 are the 10 th roots of unity, then (1 + ω) (1 + ω 2 ) (1 + ω 9 ) is A) 1 B) 1 C) 10 D) 0

K-stability and Kähler metrics, I

z, w = z 1 w 1 + z 2 w 2 z, w 2 z 2 w 2. d([z], [w]) = 2 φ : P(C 2 ) \ [1 : 0] C ; [z 1 : z 2 ] z 1 z 2 ψ : P(C 2 ) \ [0 : 1] C ; [z 1 : z 2 ] z 2 z 1

Lecture Notes 1: Vector spaces

8. Prime Factorization and Primary Decompositions

An eightfold path to E 8

Mostow Rigidity. W. Dison June 17, (a) semi-simple Lie groups with trivial centre and no compact factors and

Math 210C. A non-closed commutator subgroup

Infinitesimal Einstein Deformations. Kähler Manifolds

Notes on nilpotent orbits Computational Theory of Real Reductive Groups Workshop. Eric Sommers

A method for construction of Lie group invariants

Mic ael Flohr Representation theory of semi-simple Lie algebras: Example su(3) 6. and 20. June 2003

CW-complexes. Stephen A. Mitchell. November 1997

SPRING 2006 PRELIMINARY EXAMINATION SOLUTIONS

Orthogonal Complex Structures and Twistor Surfaces

arxiv: v3 [math.dg] 19 Jun 2017

Lecture on Equivariant Cohomology

Tropical Constructions and Lifts

Math 350 Fall 2011 Notes about inner product spaces. In this notes we state and prove some important properties of inner product spaces.

ELEMENTARY LINEAR ALGEBRA

A PRIMER ON SESQUILINEAR FORMS

AN INTEGRAL FORMULA FOR TRIPLE LINKING IN HYPERBOLIC SPACE

Cubic curves: a short survey

Linear Algebra March 16, 2019

Transcription:

SURVEYING POINTS IN THE COMPLEX PROJECTIVE PLANE LANE HUGHSTON AND SIMON SALAMON arxiv:1410.5862v2 [math.dg] 30 Sep 2015 We classify SIC-POVMs of rank one in CP 2, or equivalently sets of nine equally-spaced points in CP 2, without the assumption of group covariance. If two points are fixed, the remaining seven must lie on a pinched torus that a standard moment mapping projects to a circle in R 3. We use this approach to prove that any SIC set in CP 2 is isometric to a known solution, given by nine points lying in triples on the equators of the three 2-spheres each defined by the vanishing of one homogeneous coordinate. We set up a system of equations to describe hexagons in CP 2 with the property that any two vertices are related by a cross ratio (transition probability) of 1/4. We then symmetrize the equations, factor out by the known solutions, and compute a Gröbner basis to show that no SIC sets remain. We do find new configurations of nine points in which 27 of the 36 pairs of vertices of the configuration are equally spaced. INTRODUCTION A symmetric, informationally complete, positive-operator valued measure or SIC-POVM on the Hermitian vector space C n is a set {P j } of n 2 rank-one projection operators such that 1 n 2 P i = I, n i=1 and tr(p j P k ) = 1 n + 1 (nδ jk + 1) for all j, k. Such objects attracted wide attention following conjectures about their existence made by Zauner [43] in 1999 and Renes et al. [32] in 2004, and since then have been investigated by a large number of authors, along with higher rank versions and the allied concept of mutually unbiased basis. See, for example, [2, 3, 4, 11, 13, 17, 18, 21, 24, 33, 42, 44], and references cited therein. SIC-POVMs arise in the theory of quantum measurement (see Davies [12] and Holevo [22] for the significance of general POVMs), and are of great interest in connection with their potential applications to quantum tomography. The idea is the following. Suppose that one has a large number of independent identical copies of a quantum system (say, a large molecule), the state (or structure ) of which is unknown and needs to be determined. A SIC-POVM can be thought of as a kind of symmetrically oriented machine that can be used to make a single tomographic measurement on each independent copy of the molecule, with the property that once the results of the various measurements have been gathered for a sufficiently large number of molecules, the state of the molecule can be efficiently determined 1

2 to a high degree of accuracy. The symmetric orientation is not with respect to ordinary three-dimensional physical space (as in the classical tomography of medical imaging), but rather with respect to the space of pure quantum states. Since each element P i of a SIC-POVM is a matrix of rank one and trace unity, it determines a point in complex projective space CP n 1. It is well known that a SIC-POVM can then be defined as a configuration of n 2 points in CP n 1 that are mutually equidistant under the standard Kähler metric [29, 39]. This is the definition that we shall adopt in 3, and the distance is determined by Lemma 3.6. Such a set of points is often called a SIC, but we favour the expression SIC set. The existence of such configurations (for example, nine equidistant points in CP 2, or sixteen equidistant points in CP 3 ) is counterintuitive to our everyday way of thinking in which a regular simplex in R n has n + 1 vertices (but see [19]). It has been conjectured that CP n 1 possesses such a configuration for every n [32, 43]. There is evidence for this for n up to at least 67, and various explicit solutions have been found in lower dimensions. Most of the known SIC sets in higher dimensions are constructed as orbits of a Heisenberg group W H acting on CP n 1 (see Section 3), and representative vectors occur as eigenvectors of an isometry that is an outer automorphism of W H. In the case n = 5, the automorphisms of W H play a key role in the construction of the celebrated Horrocks-Mumford bundle over CP 4 in [23], which is an excellent reference for this group theory. In the case n = 3 (and more generally, when n is prime) any finite group of isometries whose orbit is a SIC set must be conjugate to W H [44], but in this paper we work without the assumption of group covariance (see Grassl [18]). The space CP 1, endowed with the Fubini-Study metric, is isometric to the standard two-sphere, and embedding this in R 3 is a simple example of the representation of CP n 1 as an adjoint orbit in the Lie algebra su(n) of its isometry group. The existence of a SIC set can then be interpreted as a statement about the placement of such orbits. The problem can also be formulated so as to apply to more general (co-)adjoint orbits in a Lie algebra. The vertices of any inscribed regular tetrahedron in S 2 provide a SIC set for CP 1 (n = 2). The situation for the projective plane CP 2 is already surprisingly intricate, and the case n = 3 is characterized by the existence of continuous families of noncongruent SIC sets. It is easy to begin their study. Using homogeneous coordinates, any three equally-spaced points on the equator {[0, z 2, z 3 ] : z 2 = z 3 } of the twosphere z 1 = 0 lie in a SIC set formed by adding three equally-spaced points from each of the equators of the two-spheres z 2 = 0 and z 3 = 0. If the diameter of CP 2 is chosen to be π, all nine points are a distance 2π/3 apart. Moreover, if the three

triples match up so as to lie on a total of twelve projective lines, the nine points are the flexes of a plane cubic curve [24]. In this paper, we show that any SIC set in CP 2 is congruent to one of those just described (see Theorem 5.5). This result will not surprise the experts; it has perhaps been verified numerically, and is apparently a consequence of computeraided results in [36]. Our proof relies on a computation for its final step but is predominantly analytical. We use the two-point homogeneity of CP 2 to fix two points of a SIC set; applying the moment mapping relative to a maximal torus shows that the remaining seven points lie in a pinched torus above a circle C in R 3 (illustrated in Figure 1). We exhibit the known solutions in a different form (Proposition 6.2) and characterize them by a symmetry condition (Lemma 8.2). Adding three more points a distance 2π/3 from the first two and from each other leads to a polynomial equation that is symmetric in three variables x, y, z that represent the tangents of angles measured around C (Theorem 7.3). The resulting geometry is illustrated in Section 8. Adding a sixth point allows us to write down four equations in four variables t, x, y, z. When these are totally symmetrized, we obtain a system that represents a necessary condition for the six points to form part of a SIC set. For the known solutions, at least one of the four points on the pinched torus must project to C with an angle equal to ±π/6. This fact enables us to focus attention on the so-called quotient ideal that parametrizes extra solutions, and to describe it by means of an appropriate Gröbner basis. Once one root t is fixed, the extra solutions form a finite set and the final step is to determine its size. There are too few extra solutions for these to arise from an undiscovered SIC set. This paper had its origins in a number of survey talks aimed at bringing elements of the SIC-POVM problem in various low dimensions to the attention of a wider audience, and the title and figures reflect this. We focus on the case n = 3 from Section 4 onwards, and Sections 6 10 contain the more specialized material required to achieve our goal. The Fubini-Study metric on an ambient projective space plays a central role in the construction or approximation of Kähler-Einstein metrics on algebraic varieties, and it is our hope that more general theory may shed further light on the discrete problem outlined above. 3 1. HERMITIAN PRELIMINARIES We begin with a few remarks to fix conventions. The complex vector space (1.1) C n = {z = (z 1,... z n ) : z i C}

4 of column vectors comes equipped with a Hermitian form n (1.2) w, z = w z = w i z i which is anti-linear in the first (bra) position. Each fixed w defines a linear functional z w, z, and (1.3) w w, is an anti-linear bijective mapping h: V V, equivalently an isomorphism V = V of complex vector spaces. Complex projective space is the quotient (1.4) CP n 1 = Cn \ 0 C, consisting of one-dimensional subspaces of C n or rays, and is a compact topological space. For any non-zero w C n the associated point in CP n 1 will be denoted by [w]. Each such point determines a conjugate hyperplane W defined by (1.5) W = P(ker h(w)) = CP n 2 CP n 1. This is the geometrical content of the map h. Two points [w], [z] lie on a unique projective line L = CP 1. The associated conjugate hyperplanes W, Z intersect L in [w ], [z ], where (1.6) w = w, z w w, w z, z = z, z w z, w z. The resulting four points, taken in the order [w], [z], [z ], [w ], have inhomogeneous coordinates (1.7), 0, z, z / z, w, w, z / w, w, and a real cross ratio (1.8) κ([w], [z]) = i=1 w, z z, w w, w z, z = w, z 2 [0, 1]. w 2 z 2 When the points of CP n 1 are interpreted as pure quantum states, κ can be regarded as a transition probability [7, 9, 16, 25, 26]. The Fubini-Study distance d between the points [w] and [z] is defined by expressing the cross ratio as cos 2 (d/2), so that ( ) w, z (1.9) d([w], [z]) = 2 arccos [0, π]. w z

When n = 2 we get CP 1 = S 2. We shall see in Example 2.2 that d is the spherical distance (1.10) θ = arccos u, v, u, v S 2, measuring the arclength of a great circle joining u and v. The CP 1 calculation confirms that d is the usual distance measured along geodesics of CP n 1 since any two points of the latter lie on a unique projective line CP 1. The distance (1.9) satisfies the triangle inequality (1.11) d([w], [z]) d([w], [y]) + d([y], [z]). This can be verified by working inside the CP 2 that contains [w], [y], [z]. The so-called Fubini-Study metric is the square ds 2 of the infinitesimal distance between [z] and [z + dz], computed using (1.12) κ([z], [z + dz]) = z 2 + 2 Re z, dz + z, dz 2 / z 2 z 2 + 2 Re z, dz + dz 2 = 1 dz 2 z + z, dz 2 + O( dz 3 ). 2 z 4 There are no first-order terms, and we obtain the Riemannian metric g = ds 2 where (1.13) ds 2 = 4 z 2 dz 2 z, dz 2 z 4. If we set z n = 1, and use the summation convention over the remaining indices z 1,..., z n 1, then in the traditional notation we have (1.14) g αβ dz α dz β = 4 (z αz α + 1)dz β dz β z α z β dz α dz β (z α z α + 1) 2. See, for example, Arnold [6] and Kobayashi and Nomizu [28]. When n = 2, we obtain the classical first fundamental form (1.15) ds 2 = 4 dzdz (1 + z 2 ) = 8(dx2 + dy 2 ) 2 (1 + x 2 + y 2 ) 2 on the two-sphere S 2, in which x, y are isothermal coordinates. 2. THE SPECIAL UNITARY GROUP The Hermitian form h is invariant under the action of the unitary group (2.1) U(n) = {X C n,n : XX = I}. 5

6 Its centre consists of scalar multiples e it I that act trivially on CP n 1. So we consider the special unitary group (2.2) SU(n) = {X U(n) : det X = 1}, whose centre is Z n = e 2πi/n I. The next result is due to Wigner [41]; a modern treatment is given in [15]. Theorem 2.1. The isometry group of the Fubini-Study space CP n 1, i.e. the group of bijections preserving the distance d, is generated by SU(n)/Z n and [z] [z]. The Lie algebra su(n) can (as a vector space) be defined as the tangent space T I SU(n) at the identity. It consists of tangent vectors A = Ẋ0 to curves X t = I + ta + O(t 2 ) in U(n). Thus (2.3) su(n) = {A C n,n : A + A = 0, tr A = 0}. A matrix M SU(n) acts on su(n) by the adjoint representation (2.4) A MAM 1 = MAM. The space su(n) carries an invariant inner product (2.5) A, B = tr(ab), and SU(n) itself carries a bi-invariant Riemannian invariant. We shall work with the corresponding affine space (2.6) H n = {A C n,n : A = A, tr A = 1} of Hermitian matrices of trace one. There is an obvious bijection (2.7) H n = su(n), given by A i(a n 1 I). The canonical embedding of CP n 1 into H n is a variant of the moment mapping for the adjoint action of SU(n). To describe it, assume for convenience that all vectors are normalized. Thus, we set z = 1 (z S 2n 1 ) and there remains only a phase ambiguity in passing to a point [z] = [e it z] of CP n 1. Map [z] to (2.8) P z = zz = z 1 2 z 1 z 2 z 1 z 3 z 2 z 1 z z 2 z 2 z 3 z 3 z 1,

7 which is a projection operator (meaning P 2 = P ) of rank one. The injective map (2.9) i: CP n 1 H n defined by [z] P z is SU(n)-equivariant. We can use it to measure distances since (2.10) κ([w], [z]) = w, z 2 = tr(p w P z ), assuming z = 1 = w. Moreover, the derivative (2.11) i : T x CP n 1 T x H n = R N is U(n 1)-equivariant, and (2.9) is an isometric embedding. Example 2.2 (The Bloch sphere). For n = 2, the image of this map consists of the matrices ( z1 2 ) ( ) z 1 z 2 (2.12) z 1 z 2 z 2 2 = 1 1 + a b + ic 2 b ic 1 a with z 1 2 + z 2 2 = 1 and a 2 + b 2 + c 2 = 1. This provides the well-known isomorphism CP 1 = S 2. The angle θ between two unit vectors in R 3 is given by (2.13) aa + bb + cc = cos θ. The inner product in H 2 is then 1 (2.14) 2 (1 + aa + bb + cc ) = 1 2 (1 + cos θ) = cos2 (θ/2). But θ is also the standard distance d along the great circle on the surface of the sphere joining the endpoints of the two unit vectors. In the example above, fix (say) the north pole p S 2, and consider the function κ p = sin 2 (θ/2) where θ is now latitude in radians. Its gradient κ p is tangent to the meridians joining p to the south pole p, whereas J( κ p ) is a vector field that represents rotation about pp. This situation is generalized to higher dimensions as follows. The composition (2.15) CP n 1 H n = su(n), where [z] i(p z n 1 I), is a moment mapping of the type determined whenever a Lie group acts on a symplectic manifold. The image (isomorphic to CP n 1 ) inside su(n) is an orbit for the action of SU(n). Any such adjoint orbit carries a Kähler metric by general principles. Fix a point p = [z] CP 1, and consider the function κ p defined by (2.16) κ p ([w]) = κ([z], [w]) = tr(p z P w ). We have

8 Proposition 2.3. The rotated gradient J( κ p ) is the infinitesimal isometry (Killing field) associated to i(p z n 1 I). For further details of various aspects of the Kählerian geometry of the space of pure quantum states, see Anandan and Aharonov [1], Ashtekar and Schilling [7], Bengtsson and Zyczkowski [8], Brody and Hughston [9], Gibbons [16], Hughston [25, 26], and Kibble [27]. We choose to begin with 3. SETS OF POINTS IN PROJECTIVE SPACE Definition 3.1. A SIC-POVM or SIC set is a collection S of n 2 points [z i ] in CP n 1 that are mutually equidistant, so if z i = 1 then (3.1) z i, z j 2 = κ, i j, for some fixed cross ratio κ [0, 1). We can associate to [z i ] the point P i = P [zi ] in H n. A SIC set then consists of a regular simplex embedded in (3.2) H n = su(n) = R N, N = n 2 1 with n 2 vertices {P i } that lie in the adjoint orbit CP n 1. The latter requirement is the crucial one, since a regular simplex with n 2 vertices in R N is readily obtained by projecting an arbitrary orthonormal basis of R N+1. Do SIC sets exist? Example 3.2. A SIC set in CP 1 = S 2 is an inscribed tetrahedron in the two-sphere. Any two such tetrahedrons are congruent by SO(3) = SU(2)/Z 2, though that does not stop us seeking the neatest set of vertices to write down. One set is { (3.3) [0, 1], [ 2, 1], [ 2, ω], [ } 2, ω 2 ], where ω = e 2πi/3. Another set of vertices, which is perhaps less obvious, is { } (3.4) [1, ϖ], [ϖ, 1], [1, ϖ], [ϖ, 1], where ϖ = (1 + i)/(1 + 3). This second set nevertheless plays an important role, as we shall see. If n 3, any two SIC sets in CP n 1 R N N = n 2 1), but not in general by SU(n). are congruent by SO(N) (where

One can present more SIC sets by generalizing the second tetrahedron (3.4). We define two cyclic groups of order n. Let W be the group generated by the cyclic permutation (3.5) [z 1, z 2,..., z n ] [z n, z 1,..., z n 1 ]; let ω = e 2πi/n, and denote by H the group generated by (3.6) [z 1, z 2,..., z n ] [z 1, ωz 2,..., ω n 1 z n ]. W H acts on CP n 1 as a subgroup of SU(n) isomorphic to Z n Z n. This subgroup is sometimes called the Weyl-Heisenberg group after [40]. It can be regarded as the projectivization of an extended finite group, namely the Heisenberg group of three-by-three matrices with coefficients in the ring Z n. For this reason, it is legitimate to refer to the action of W H simply as that of the Heisenberg group. The following two results can be verified by direct calculation: Proposition 3.3. The orbit (3.7) (W H) [0, 1, 1] is a SIC set consisting of nine points in CP 2. Proposition 3.4. Let r = 2 and s = 2 + 5. Then (3.8) (W H) [ s i(r+s), 1 r + i, s + i(s r), 1 + r + i] is a SIC set of sixteen points in CP 3. An element z C n such that the orbit (W H) [z] is a SIC set is called a fiducial vector for the action of W H. In his 1999 Vienna PhD thesis [43], Zauner made a number of conjectures that extended the basic Conjecture 3.5. CP n 1 possesses a SIC set for all n. It is widely believed that such a set can always be realized as an orbit of W H, and that the number of non-congruent solutions (meaning solutions that are not related to one another by an isometry or element of SU(n)) increases with n. There are sporadic constructions of SIC sets using different finite groups (see Remark 3.8). Explicit algebraic solutions are known for n = 2, 3, 4,..., 15, 19, 24, 35 and 48, from work of Zauner [43], Appleby [2], Renes et al. [32], Flammia [14], Grassl [18], Zhu [44], and many other authors (see [3, 11] and references cited therein). All such examples lie (up to isometry) in solvable extensions of Q [4]. Extensive numerical verification has been carried out for n 67 (Scott and Grassl [33]). 9

10 The next result is well known, but we include it for completeness. Let {[z i ]} be a SIC set in CP n 1 and {P i } its image in H n. Recall that tr(p i P j ) = κ if i j. Thus κ is the cross ratio or transition probability between any two points in the SIC set. Lemma 3.6. Any SIC set in CP n 1 satisfies κ = 1/(n + 1), and 1 (3.9) Pi = I. n Proof. Define Q j = P j κi. Then { 1 κ i = j (3.10) tr(p i Q j ) = 0 i j So (P i ) is a basis of iu(n) (called a quorum) and we can set n 2 (3.11) I = c i P i. i=1 Applying tr(q j ), we get 1 κn = c j (1 κ), so all the c i are equal. To complete the proof, take the trace of (3.11). This gives (3.12) n = n 21 κn 1 κ, and κ = (1 n)/(1 n 2 ) = 1/(1 + n). It will be convenient in our analysis of SIC sets to introduce the following. Definition 3.7. Two points in CP n 1 will be said to be correctly separated if the cross ratio that they define equals 1/(n + 1). Suppose that CP n 1 admits a SIC set S. Then, up to isometry, two points form part of a SIC set if and only if they are correctly separated. This follows from the fact that CP n 1 is a two-point homogeneous space, meaning that there exists an isometry that maps any two points to any other two points the same distance apart [38]. The lemma above is then a key result that enables one to go some way in attempting to construct a SIC set without knowing for sure that it exists. Remark 3.8. Lemma 3.6 precludes the existence of four or more points of a SIC set from lying on a projective line CP 1 whenever n 3, since their cross ratio would have to be that for n = 2, namely 1/3. An application relates to the SIC set in CP 7 constructed by Hoggar [21]. It consists of a (Z 2 ) 6 orbit of 64 points that the Hopf fibration π : CP 7 HP 3 projects down to an equal number of points in the quaternionic projective space HP 3. It would be impossible to find a SIC set in CP 7

with four points in each fibre of π, but we wonder whether there exists a SIC set arising from 32 points in HP 3 with two points in each fibre. Such questions are related to work by Armstrong et al. on twistor lifts [5]. 4. THE ACTION OF A MAXIMAL TORUS Starting in this section, we restrict the discussion mainly to the case n = 3. We shall develop the concept of moment mapping, but restricted to a maximal torus in SU(3), acting on CP 2. Fix the torus (4.1) T = eix 1 0 0 3 0 e ix 2 0 : x i = 0 mod 2π 0 0 e ix 3 i=1, which is, of course, homeomorphic to S 1 S 1. The hyperplane x 1 + x 2 + x 3 = 0 in R 3 represents the Lie algebra t of T, which we also identify with t using the induced inner product. The moment mapping for T acting on CP 2 is then the composition (4.2) CP 2 su(3) t obtained by projecting the adjoint orbit orthogonally to t. When we pass from su(3) to H 3 via (2.7), we can identify this composition with the mapping [z] (x 1, x 2, x 3 ), where (4.3) (x 1, x 2, x 3 ) = µ([z]) = 1 z 2( z 1 2, z 2 2, z 3 2 ) consists of the diagonal entries in (2.8). Here, z 2 = z 1 2 + z 2 2 + z 3 2, though it is convenient to assume z = 1. After the shift from traceless matrices to H 3, the image of µ is the two-simplex T, a filled equilateral triangle lying in the plane x 1 + x 2 + x 3 = 1, illustrated in Figure 1. The residual three-fold symmetry visible is that of the Weyl group W = N(T )/T = Z 3. It is well known that T parametrizes the orbits of T on CP 2 via (4.3). See, for example, Guillemin and Sternberg [20]. The inverse image of an interior point of T is a two-torus T/Z 3 ; the inverse image of a vertex is a single point in CP 2 ; and the inverse image of any other boundary point is a circle S 1. Topologically, this leads to a description of the complex projective plane as a quotient (4.4) CP 2 = T T 2. Here is the equivalence relation that collapses points over the boundary of T in accordance with the scheme outlined above. 11

12 Let m 1, m 2, m 3 be the midpoints of the sides of T, and consider the circles (4.5) C i = µ 1 (m i ), i = 1, 2, 3. The first circle C 1 consists of those points [0, z 2, z 3 ] of CP 2 with z 2 = z 3. Any set of three equidistant points in C 1 has the form (4.6) [0, e iσ, 1], [0, e iσ, ω], [0, e iσ, ω 2 ], where ω = e 2πi/3. The cross ratio defined by any two of these points is given by (4.7) 1 2 (1 + ω) 2 = 1 4, so they are indeed correctly separated. Similarly, C 2 consists of points [z 1, 0, z 3 ] with z 1 = z 3, and C 3 points [z 1, z 2, 0] with z 1 = z 2. Now choose three equidistant points in C 2, and three equidistant ones in C 3. It is easy to check that the resulting nine points constitutes a SIC set. This generalizes Proposition 3.3. Definition 4.1. By a midpoint solution, we mean a SIC set in CP 2 consisting of three points in each of the three circles C 1, C 2, C 3. FIGURE 1. The image of the moment mapping µ: CP 2 R 3 is a two-simplex T that takes the form of a filled equilateral triangle. If z = (z 1, z 2, z 3 ) is a unit vector in C 3, the point [z] is mapped to (x 1, x 2, x 3 ) = ( z 1 2, z 2 2, z 3 2 ). The inscribed circle is the intersection of the plane x 1 + x 2 + x 3 = 1 containing the (coloured) image of µ and the (invisible) sphere x 2 1 + x 2 2 + x 2 3 = 1 2.

This construction defines a one-parameter family of SIC sets up to isometry, since the stabilizer in SU(3) of the points in C 1 is a subgroup U(1) that can be used to remove the phase ambiguity in C 2. See the discussion surrounding (6.10). Let C denote the circle passing through the midpoints m 1, m 2, m 3, illustrated in Figure 1. As a curve in R 3, it is the intersection of the plane x 1 + x 2 + x 3 = 1 with the sphere x 2 1 + x 2 2 + x 2 3 = 1/2. It was actually plotted using the next result. Lemma 4.2. In R 3, the inscribed circle C is parametrized by (4.8) (cos 2 θ, cos 2 (θ + 2 3 π), cos2 (θ + 4 ) 3 π), for θ [ 1 2 π, 1 2 π). 2 3 Proof. First, consider the effect of µ on real vectors. Suppose that z = (a, b, c) is a unit vector with a, b, c R, and set s k = a k + b k + c k. There is an identity (4.9) s 1 ( a + b + c)(a b + c)(a + b c) = s 2 2 2s 4. If µ([z]) = (a 2, b 2, c 2 ) C then s 2 = 1 and s 4 = 1/2, so the right-hand side of (4.9) vanishes. It follows a ± b ± c = 0 for some choice of signs. Now set 2 (4.10) a = 3 cos θ, b = 2 3 cos(θ + 2 3 π), c = 2 3 cos(θ + 4 3 π). Trig-expanding b and c shows that a + b + c = 0 and a 2 + b 2 + c 2 = 1. It follows from (4.9) that s 4 = 1/2 and (a 2, b 2, c 2 ) C. The midpoints m 1, m 2, m 3 are given respectively by θ = ±π/2, π/6, π/6. This confirms the stated range for t. 5. WEYL-HEISENBERG ORBITS In this section, we show how the moment mapping (4.3) helps one to understand the action of the groups W and H defined in (3.5) and (3.6) with n = 3. We shall see that the C plays a prominent role, and Lemma 4.2 will be the basis for the parametrization of elements of a SIC set. Lemma 5.1. The H -orbit of a point [z] in CP 2 consists of three points that are correctly separated from one another if and only if µ([z]) C. Proof. Suppose that z = z (0) is a unit vector. The orbit H [z] consists of the projective classes of the vectors (5.1) z (0) = (z 1, z 2, z 3 ), z (1) = (z 1, ωz 2, ω 2 z 3 ), z (2) = (z 1, ω 2 z 2, ωz 3 ) generated by (3.6). We can express (5.2) z (0), z (1) 2 = ( z 1 2 + ω z 2 2 + ω 2 z 3 2 )( z 1 2 + ω 2 z 2 2 + ω z 2 3 ) 13

14 in the form α β, where (5.3) α = z 1 4 + z 2 4 + z 3 4, β = z 2 2 z 3 2 + z 3 2 z 1 2 + z 1 2 z 2 2. Therefore z (0) and z (1) are correctly separated if and only if α β = 1/4. But (5.4) α + 2β = ( z 1 2 + z 2 2 + z 3 2 ) 2 = 1, since z = z (0) is normalized, so the condition of correct separation is α = 1/2. Since x i = z i 2 are the Cartesian coordinates in R 3, correct separation of z (0) and z (1) implies that µ([z]) C. This condition only depends on µ([z]) since H is a subgroup of T and its action commutes with all the elements of T. Therefore if µ([z]) C, all three points in (5.1) will be correctly separated. Example 5.2. Lemma 5.1 is really an assertion about the induced metric on the fibres µ 1 (p) for p C. This metric will depend crucially on the position of p in C, since it degenerates as p approaches any one of the midpoints m i (over which the fibres are circles rather than 2-tori). This behaviour is illustrated in Figure 2, which provides a visualization of the fibres µ 1 (p i ) for i = 1, 2, where (5.5) p 1 = ( 2 3, 1 6, 1 6 ), p2 = ( 1 8 (3 + 5), 1 4, 1 8 (3 5) ). are two points of C. Note that p 1 is the point diametrically opposite m 1, whereas p 2 lies between p 1 and m 3. The coordinates used in Figure 2 are derived from the action of the maximal torus (4.1), which is represented by translation. Scalar multiplication by ω = e 2πi/3 on vectors in C 3 generates the action of the centre Z 3 of SU(3), so that (z 1, z 2, z 3 ) and (ωz 1, ωz 2, ωz 3 ) appear as distinct points in the diagrams, although they determine the same point of CP 2. The centre is responsible for the evident three-fold symmetry, which is best represented by the hexagonal fundamental domain on the right-hand side. Comparing this with the left-hand parallogram and its translates, one sees that a 2-torus can be formed by identifying the opposite edges of a hexagon, a fact that is well known (see, for example, Thurston [37]). Both diagrams display exactly three distinct points of CP 2 in the closure of each coloured fundamental domain, and each of these triples of points forms an equilateral triangle. This can be seen from an inspection of the curves that are the loci of points a distance 2π/3 from the centre point. The latter is correctly separated from each of the other two points, and these two points are correctly separated from each other because distances are translation invariant. We are now in a position to give a full description of those SIC sets that are orbits of the group generated by (3.5) and (3.6).

15 FIGURE 2. A representation of the fibre µ 1 (p 1 ) (left) and µ 1 (p 2 ) (right) in CP 2 for the points (5.5). The coloured regions are two different fundamental regions for the torus, and the red curves are points a distance 2π/3 from the centre point. Theorem 5.3. Let z = (z 1, z 2, z 3 ) C 3. Then (W H) [z] is a SIC set if and only if one of the variables z 1, z 2, z 3 vanishes, or (5.6) [z] = [cos θ, ω j cos(θ + 2 3 π), ωk cos(θ + 4 ] 3 π), for some θ [ 1 2 π, 1 2π) and j, k {0, 1, 2}. Proof. Suppose that z is a unit vector and that (W H) [z] is a SIC set. Let z = (z 3, z 1, z 2 ). In the notation (5.1), we have (5.7) z (k), z 2 = z1 z 3 + ω k z 2 z 1 + ω 2k z 3 z 2 2 = β + 2 Re [ ω 2k ], where (5.8) = z 2 1z 2 z 3 + z 2 2z 3 z 1 + z 2 3z 1 z 2. By assumption, (5.7) equals 1/4 for all k = 0, 1, 2. From (5.4) we have β = 1/4, so the expression in square brackets above must be purely imaginary. This happens for all k if and only if = 0. By assumption, µ([z]) C, so [z] must lie in a T -orbit of [a, b, c] where a, b, c are given by (4.10) for some θ. Since a + b + c = 0, (5.8) and (5.7) tell us that z = (a, b, c) is a fiducial vector. Let us look for other fiducials in the same T -orbit

16 by considering (5.9) (z 1, z 2, z 3 ) = (a, e iβ b, e iγ c), having normalized the coefficient of a. Let us assume that abc 0, so b + c 0. Since = 0, we have (5.10) e 3iβ b + e 3iγ c = b + c. Taking the moduli of both sides gives cos(3β 3γ) = 1, so β equals γ mod 2π/3. It follows that both β and γ are multiples of 2π/3, and that [z] has the form (5.6). Conversely, the vector (5.6) satisfies (5.7) and projects to C. Thus its W H orbit is a SIC set. To summarize, any three equally-spaced points on C form the base of a group covariant SIC set. If these are the three midpoints m i of the sides then any point in µ 1 (m i ) is a fiducial vector. But for a generic point p C, the choices are restricted to nine points on the two-torus µ 1 (p). As p approaches a midpoint, these nine points become three. Remark 5.4. The methods of this section can be extended to the study of SIC sets in CP n 1 that arise as orbits of W H for n > 3. Using the moment mapping µ: CP n 1 R n, one can define a subset of the simplex µ(cp n 1 ) consisting of points whose inverse image contains H -orbits of correctly-separated points. For CP 3 the relevant subset consists of two circular arcs inside a solid tetrahedron, but is no longer one-dimensional if n > 4, as discussed by Lora Lamia [30]. For applications of the use of µ: CP 3 R 4 in classifying almost-hermitian structures on manifolds of real dimension six, see Mihaylov [31]. The SIC sets in CP 2 described above have been discussed by Renes et al. [32], Zhu [44], and various other authors. In particular, it is known that any SIC set arising from Theorem 5.3 is isometric to a midpoint solution. This can be proved by adapting the proof of Proposition 6.2 below, but we shall prove a much stronger result in this paper, namely Theorem 5.5. Any SIC set S in CP 2 is congruent modulo SU(3) to a midpoint solution. In the next section, we shall work with yet another description of the isometry class of a midpoint solution, in which each circle C i contains exactly two points of the SIC set.

17 6. TWO-POINT HOMOGENEITY Suppose, going forward, that S is a SIC set in CP 2, consisting of nine points [z i ], i = 1,..., 9. Up to the action of the isometry group, we are free to assume that S contains the two points of C 1 represented by the unit vectors (6.1) z 1 = 1 2 (0, 1, ω), z 2 = 1 2 (0, 1, ω 2 ), which are a distance 2π/3 apart. This is on account of the two-point homogeneity of CP 2. Lemma 3.6 tells us that any other point [z] of S must satisfy (6.2) z, z j 2 = 1 4 z 2, j = 1, 2. Using this equation, we can prove another lemma that emphasizes the important role played by the incircle C. Lemma 6.1. The moment map µ projects any remaining point [z] of S to a point of C. Indeed, we may take z to be a unit vector of the form ( (6.3) z(σ, φ) = 2 3 e iσ cos φ, cos(φ + 2 3 π), cos(φ + 4 ) 3 π) for some σ ( π, π] and some φ ( 1 2 π, 1 2 π]. To lighten the notation, we shall write z[σ, φ] as a shorthand for [z(σ, φ)], so that square brackets on either side of z indicate a projective class. Lemma 4.2 tells us that z[σ, φ] lies over C. Observe that µ(z[σ, φ]) depends only on the angle φ measured around C, and not on the phase σ. Moreover, as σ and φ vary, z[σ, φ] parametrizes a pinched two-torus, the pinch point being (6.4) z[σ, 1 2 π] = z[σ, 1 2π] = [0, 1, 1], which is evidently independent of σ. Having chosen [z 1 ], [z 2 ] on C 1, we can see that any third point of C 1 S must be this point, which explains the pinching. One should note that z(σ, φ) = z(σ, φ + π), which is why φ = π/2 is excluded from the non-projective representation (6.3). Proof of Lemma 6.1. Let us suppose that [z] = [z 1, z 2, z 3 ] S. If z 2 = 0 then [z] C 2 and (6.3) will be valid for φ = π/6. We may therefore take z 2 = 1 and set z 1 = a, z 3 = c where a, c C. Then by assumption, we have (6.5) 1 cω 2 = 1 cω 2, which implies that c is real, and (6.6) [z] = [a, 1, ± c ].

18 Using (6.2), we see that (6.7) so a 2 = (1 ± c ) 2, and 1 2 (1 ± c + c 2 ) = 1 4 ( a 2 + 1 + c 2 ), (6.8) a 2 + 1 + c 2 = 2(1 ± c + c 2 ). Therefore, (6.9) a 4 + 1 + c 4 = 2 ± 4 c + 6 c 2 ± 4 c 3 + 2 c 4 = 1 2 ( a 2 + 1 + c 2 ) 2. It follows that µ does indeed map [z] into C. In view of Lemma 4.2, we must be able to express [z] in the stated form for some e iσ U(1). The points [z 1 ], [z 2 ] are both fixed by the subgroup U(1) of (4.1) generated by (6.10) e 2ix 0 0 0 e ix 0, 0 0 e ix so we may assume that a third point of S is z[0, θ]. The next result shows that there does exist a SIC set containing this point for any θ. Proposition 6.2. For any θ [ 1 2 π, 1 2π), the six points [0, 1, ω] = [z 1 ], [0, 1, ω 2 ] = [z 2 ] C 1, (6.11) combine with three points [1, 0, ω] = z[ 2 3 π, 1 6 π], [1, 0, ω2 ] = z[ 2 3 π, 1 6 π] C 2, [1, ω, 0] = z[ 2 3 π, 1 6 π], [1, ω2, 0] = z[ 2 3 π, 1 6 π] C 3, (6.12) z[0, θ 1 3 π], z[0, θ], z[0, θ + 1 3 π] to form a SIC set isometric to a midpoint solution. We shall denote this SIC set by S θ ; it is illustrated in Figure 3. Proof. Consider the matrix (6.13) M = 1 ω2 ω 1 1 ω ω 2 3 1 1 1 It is easy to check that M U(3) and that M 3 = iω 2 I. A calculation shows that. (6.14) M z(0, θ) = 1 2 ( e iθ ω 2, e iθ, 0 ),

so that M maps z[0, θ] to the point [e 2iθ, ω, 0] of C 3. Moreover, M maps the array (6.11) to the array (6.15) [0, 1, 1], [1, 0, ω], [1, 0, ω 2 ], [0, 1, ω 2 ], [0, 1, ω], [1, 0, 1] of points in C 1 C 2. It follows that M maps S θ onto three triples of points, each triple belonging to C i for some i = 1, 2, 3. The first six points (6.11) of S θ do not depend on θ, whereas the last triple of points can be rotated at will (by varying θ) around a circle C 3 covering C. For example, z[0, 0] lies over the point p 1 of C diametrically opposite m 1 (see (5.5)). Remark 6.3. Nine points in CP 2 are the inflection points of a non-singular cubic curve if and only if the line determined by any two of them contains a third. This being the case, there are twelve such lines altogether, on which the nine points lie by threes, with four of the twelve lines through each of the nine points, thus forming the so-called Hesse configuration {9 4, 12 3 }. For the points of S θ to arise in this way, and as described by Hughston [24] and Dang et al. [11], the projective line L 1 = CP 1 generated [z 1 ], [z 2 ] must contain a third point of S θ. But L 1 is the inverse image by µ of the side of T containing m 1, and will only contain another point if 19 FIGURE 3. The SIC set S θ defined by Proposition 6.2 contains two points in each circle C i (including [z 1 ], [z 2 ] in C 1 ) that do not depend on θ, together with a triple of points (including [z 3 ]) that µ also projects to C for which θ represents the angle around C.

20 θ assumes one of the values ±π/2, ±π/6. This occurs when the three red legs (the ones generated by [z 3 ] by rotation by 2π/3) in Figure 3 line up with the green legs (the ones over the midpoints), and S θ is then itself a special midpoint solution. Example 6.4. The unitary transformation M maps C 3 to C 3. It permutes the elements of the SIC set (W H) [0, 1, 1], though it fixes none of them. The matrices (6.16) A = 0 1 0 0 0 1, B = 1 0 0 0 ω 0 1 0 0 0 0 ω 2 generate W and H respectively, and satisfy (6.17) MAM 1 = ωb, MBM 1 = ω 2 A 1 B 1. It follows that M is an element of the so-called Clifford group, the normalizer of W H in U(3). Modulo phase, this normalizer is isomorphic to a semidirect product SL(2, Z 3 ) (Z 3 ) 2 (see Appleby [2] and Horrocks-Mumford [23]). Equation (6.17) asserts that M induces the automorphism of W H given by ( ) 0 1 (6.18) SL(2, Z 1 1 3 ). It is conjectured that a fiducial vector can always be found in an eigenspace of some element of the Clifford group (see Zauner [43]). In the case of M, a computation shows that any one of its eigenvectors defines a point of CP 2 whose orbit under W H is a configuration of nine points arranged in nine lines. Each of the 27 pairs of points lying on one of the nine lines has a cross ratio κ = 1/3, whereas the remaining nine pairs of points have κ = 0. Compared to the Hesse configuration above, this means that three of the twelve triples of points are not collinear, but each of these three triples forms an orthonormal basis of C 3. Remark 6.5. If an isometry is to fix both [z 1 ] and [z 2 ], there is no ambiguity remaining in the choice of φ [ π/2, π/2) in Lemma 6.1. However, we are at liberty to interchange [z 1 ] and [z 2 ] by applying either complex conjugation or the unitary (6.19) 1 0 0 0 0 1 0 1 0 The former has the effect of replacing σ by σ, and the latter of replacing φ by φ in (6.3). In particular, the congruence class of the unordered set S θ uniquely specifies θ. This fact can also be verified using a triple product that measures the signed area of the planar geodesic triangle spanned by three points. See, for example, Brody and Hughston [9] and references cited therein..

21 7. TRIGONOMETRY From now on, we shall assume that S is a SIC set in CP 2 that contains the points [z 1 ], [z 2 ] defined by (6.1). Lemma 6.1 tells that any other point of S has the form [ (7.1) z[σ, φ] = e iσ cos φ, cos(φ + 2 3 π), cos(φ + 4 ] 3 π)ω2, where (σ, φ) belongs to the rectangle (7.2) R = ( π, π] ( 1 2 π, 1 2 π]. The next result, from which many others follow, translates distance into the new rectangular coordinates. Lemma 7.1. Suppose that φ, ψ ( 1 2 π, 1 2 π) \ { 1 6 π, 1 6π}. Then the points z[σ, φ] and z[τ, ψ] are the correct distance 2π/3 apart if and only if (7.3) 1 cos(σ τ) = 9(1 + 2 cos(2(φ ψ))) sec φ sec ψ. 16(cos φ cos ψ + 3 sin φ sin ψ) Proof. Not only do we have to establish the formula, but we also need to show that the assumptions imply that the denominator of the fraction is non-zero. We use the abbreviated notation (7.4) Γ 0 = cos φ cos ψ, Γ 1 Γ 2 = cos(φ + 2 3 π) cos(ψ + 2 3 π), = cos(φ + 4 3 π) cos(ψ + 4 3 π). The condition on the cross ratio for correct separation is that 4 (7.5) e i(σ τ) Γ 0 + Γ 1 + Γ 2 2 = 1 which gives 9 (7.6) 2 cos(σ τ)γ 0 (Γ 1 + Γ 2 ) + Γ 2 0 + (Γ 1 + Γ 2 ) 2 = 9 16, and therefore (7.7) 32Γ 0 (Γ 1 + Γ 2 )(1 cos(σ τ)) = 32Γ 0 (Γ 1 + Γ 2 ) + 16Γ 2 0 + 16(Γ 1 + Γ 2 ) 2 9. A calculation shows that the right-hand side of (7.7) is equal to (7.8) 9(1 + 2 cos(2(φ ψ))), which vanishes when cos(φ ψ) = ±1/2. By hypothesis, Γ 0 0. If (7.9) Γ 1 + Γ 2 = 1 2 [cos φ cos ψ + 3 sin φ sin ψ] 4,

22 vanishes, then (7.10) cos φ cos ψ + sin φ sin ψ = cos(φ ψ) = ± 1 2, and hence (7.11) cos φ cos ψ = ± 3 4, sin φ sin ψ = 1 4. Now set x = tan φ and y = tan ψ. Then xy = 1/3 and it holds that (7.12) ± 3 = tan(φ ψ) = x y 1 + xy = 3 2 (x y). We therefore have (7.13) (x + y) 2 = (x y) 2 + 4xy = 0, and φ = ψ = ±π/6, values that are excluded. We may therefore assume that Γ 0 (Γ 1 + Γ 2 ) 0, and (7.3) follows. Lemma 7.2. If S contains the pinch point [0, 1, 1] as well as [0, 1, ω] and [0, 1, ω 2 ], then S is a midpoint solution. Proof. By hypothesis, S contains three points of the circle C 1. If z[σ, φ] is a fourth point of S, then (7.1) is correctly separated from [0, 1, 1] and (7.14) cos(φ + 2 3 π) cos(φ + 4 3 π) = ± 3 2. This implies that sin φ = ±1/2, and forces z[σ, φ] to lie on C 2 C 3. Therefore S lies in the disjoint union C 1 C 2 C 3. One can rewrite (7.3) as (7.15) 9(1 + 2 cos 2(φ ψ)) cos(σ τ) = 1 4(4 cos 2 φ cos 2 ψ + 3 sin 2φ sin 2ψ) 5 + 4 cos 2φ + 4 cos 2ψ 14 cos 2φ cos 2ψ 6 sin 2φ sin 2ψ =. 4(1 + cos 2φ + cos 2ψ + cos 2φ cos 2ψ + 3 sin 2φ sin 2ψ) We shall convert the right-hand side into a rational function by setting (7.16) x = tan φ, y = tan ψ. In the light of Lemma 7.2, we assume from now on that x, y are finite. Equation (7.15) simplifies to (7.17) cos(σ τ) = 11 + 9x2 + 9y 2 27x 2 y 2 24xy. 16(1 + 3xy)

If 1 + 3xy = 0, then the numerator on top of it must also vanish, so x 2 + y 2 = 2/3 and (x + y) 2 = 0. Thus (as in the previous proof) x = y = ±1/ 3. This means that z[σ, φ] lies on one of the circles C 2, C 3, and z[τ, ψ] lies on the other, so there are no restrictions on σ and τ. The main result of this section is the following, which establishes a criterion for the existence in CP 2 of five points that are correctly separated from one another. Theorem 7.3. Suppose that z[σ, φ], z[τ, ψ], z[υ, χ] are three points of CP 2, a distance 2π/3 away from each other (and from [z 1 ], [z 2 ]). Set p = x + y + z, q = yz + zx + xy, r = xyz, where x = tan φ, y = tan ψ, z = tan χ. Then F (p, q, r) = 0 where 23 (7.18) F (p, q, r) = 9 22p 2 + 9p 4 + 87q 126p 2 q + 27p 4 q + 298q 2 226p 2 q 2 +24p 4 q 2 + 414q 3 138p 2 q 3 + 189q 4 + 27q 5 3pr 50p 3 r 15p 5 r +88pqr 48p 3 qr + 234pq 2 r + 18p 3 q 2 r 144pq 3 r + 81pq 4 r + 189r 2 480p 2 r 2 153p 4 r 2 + 1398qr 2 306p 2 qr 2 + 2736q 2 r 2 486p 2 q 2 r 2 +810q 3 r 2 + 243q 4 r 2 558pr 3 486p 3 r 3 + 2376pqr 3 810pq 2 r 3 +567r 4 162p 2 r 4 + 6399qr 4 + 486q 2 r 4 + 1701pr 5 + 2187r 6. Proof. Although (7.18) is rather complicated, the existence of such an expression is a consequence of the elementary trigonometric identity (7.19) A + B + C = 0 cos 2 A + cos 2 B + cos 2 C = 1 + 2 cos A cos B cos C, which is tailor made for (7.17). The identity itself can be proved by writing applying more standard ones to the sum A + (B + C). Denote the right-hand side of (7.17) by the symmetric function c(x, y). Then (7.20) c(x, y) 2 + c(y, z) 2 + c(z, x) 2 = 1 + 2c(x, y)c(y, z)c(z, x). This simplifes into the vanishing of the quotient (7.21) 243 f(x, y, z) 2048(1 + 3xy) 2 (1 + 3yz) 2 (1 + 3zx) 2, in which f is a totally symmetric polynomial. We can then use the Mathematica command SymmetricReduction to express (7.22) f(x, y, z) = F (p, q, r) as a function of the elementary symmetric polynomials, and the result follows.

24 8. GRAPHICAL INTERPRETATION Suppose once again that S is a SIC set in CP 2 containing [z 1 ] = [0, 1, ω] and [z 2 ] = [0, 1, ω 2 ], and (in view of Lemma 7.2) not the third point [0, 1, 1] of C 1. The planar parametrization (7.3) of the remaining points of S enables us to describe graphically the quest for such SIC sets. Before we do this, we prove two results that help with their classification. Setting φ = π/6 in (6.3) defines the circle C 2, and φ = π/6 the circle C 3. It will be convenient to consider three more circles C, C 0, C + given by σ = 2π/3, 0, 2π/3 respectively. Unlike C 1, C 2, C 3, these three are not disjoint: they meet in [0, 1, 1]. The circles C 2, C 3 are represented by horizontal lines in R, and C, C 0, C + by equally-spaced vertical lines; all five have diameter π. The lines representing C, C 0 and C 2 are visible in Figure 7. Lemma 8.1. If S contains a point z[σ, φ] with φ = π/6 then S is isometric to a midpoint solution. Proof. We can use the isometry (6.10) to shift all points of S by a translation parallel to the horizontal axis within our rectangle R. We may therefore assume that σ = 0. Suppose for definiteness that φ = π/6, so that x = 1/ 3 and z[σ, φ] C 3. Suppose that z[τ, ψ] is a fourth point of S, and apply (7.17). The numerator equals (8.1) 11 + 9x 2 + 9y 2 27x 2 y 2 24xy = 8(1 + 3y), and the right-hand side of (7.17) becomes 1/2 unless y = 1/ 3. It follows that τ = ±2π/3 or ψ = π/6. Indeed, the set of points correctly separated from [z 1 ], [z 2 ] and z[0, π/6] is the union C C 2 C +. This union must now contain six points of S, and no circle can contain more than three. Now suppose that S contains distinct points z[ 2 3 π, ψ] C + and z[υ, 1 6 π] C 2. Then (7.17) tells us that either y = 1/ 3 (and so ψ = π/6), or else (8.2) cos( 2 3 π υ) = 1 2 and υ = 2π/3 or υ = 0. So either the first point lies on C 3, or else the second point lies on C C 0. Now suppose that S contains z[ 2 3 π, ψ] C and z[ 2 3 π, χ] C + This time, (7.17) yields (8.3) (3y 2 1)(3z 2 1) = 0, and at least one of the two points is one of the last four in (6.11). We may also suppose that [0, 1, 1] S by Lemma 7.2. It then follows that S consists of [z 1 ] and [z 2 ], the two points in C (C 2 C 3 ), the two points in C 0 (C 2 C 3 ) and three points in C +, or the same thing with C and C + interchanged. Applying

(6.10) with x = ±2π/3, we obtain exactly the SIC set S θ for some θ (like the one that includes the green points in Figures 6 and 7). Then the result follows from Proposition 6.2. Lemma 8.2. Suppose that S is a SIC set that contains [z 1 ], [z 2 ] and z[0, θ]. Recall that any SIC set has this property up to isometry. If S contains distinct points z[σ 1, φ], z[σ 2, φ] with σ 1, σ 2 ( π, π) then it is isometric to a midpoint solution. Proof. First observe that σ 1 + σ 2 = 0; this follows by applying Lemma 7.1 in which we can set (τ, ψ) = (0, θ) to obtain cos σ 1 = cos σ 2. So take σ = σ 1. In view of Lemma 8.1, we may suppose that x = tan φ is different from ±1/ 3. We can choose a sixth point z[τ, ψ] of S such that τ π, since the circle τ = π can contain at most three points a distance 2π/3 apart. It follows from (7.17) that either 1 + 3xy = 0 (and we can apply Lemma 8.1) or (8.4) cos(σ τ) = cos( σ τ). Since σ = 0 and σ = π do not yield distinct points, the only possibility remaining from our assumption is that τ = 0. If t = tan θ and y = tan ψ, (7.17) implies that (8.5) t 2 + y 2 3t 2 y 2 8ty 3 = 0. This gives (8.6) y = t ± 3 1 3 t, ψ = θ ± π/3 modulo π. This is the configuration of three points visible on the central vertical axis in Figure 7. All together, S now contains at most seven points including [z 1 ] and [z 2 ], which is a contradiction. Using (7.17), one can in fact show that given the sixth point, either φ or ψ must equal ±π/6. We are now in a position to illustrate the problem of finding SIC sets that contain [z 1 ] and [z 2 ]. We can (and shall) assume that a third point of S is z[0, θ] for some fixed θ ( π/2, π/2). This point corresponds to one on the central vertical axis of the rectangle R, and will be displayed by a black dot in the figures. We shall draw some curves to illustrate the concept of correctly separated points in R, meaning that the distance between the points they represent in CP 2 equals 2π/3. A fourth point z[σ, φ] of S will be displayed by a red dot. In Figure 4, θ = π/16 so that the third point z[0, θ] of S is close to centre of R. The black curve is the set of points z[σ, φ] which are a distance 2π/3 from z[0, θ]. The remaining six points of S must therefore lie on this curve. One such example 25

26 is represented by the red dot, which actually has φ = π/4. Points z[τ, ψ] a distance 2π/3 apart from this red point are those on the red curve (which has two components). The intersection of the black and red curves consists of points which are correctly separated from both the third and fourth points. Since there are only four of these (we require five), the value x = tan φ = 1 cannot in fact occur when θ = π/16. FIGURE 4. The black (resp., red) point is correctly separated from all points on the black (resp., red) curve. The two points cannot belong to a SIC set because there are only four remaining points correctly separated from them both. FIGURE 5. The points that are correctly separated from the black point can form a disconnected set. Here, there appear to be five points correctly separated from the red and black points, but these five points do not in fact form part of a SIC set.

27 FIGURE 6. Here the fourth point z[0, 1 16 π 1 3 π] belongs to S θ which is generated by the remaining five points on the intersection of the black and red curves. FIGURE 7. Here the fourth point z[ 2 3 π, 1 6 π] S θ is equidistant from all points on the circles C, C 0, C 2, represented by straight lines in R. The segments top and bottom collapse to the pinch point. The nature of the black curve is heavily dependent on the value chosen of θ and t = tan θ. If z[π, φ] is correctly separated from z[0, θ] and x = tan φ then (8.7) 9(1 3t 2 )x 2 + 24tx + 9t 2 + 5 = 0. Computing the roots of the discriminant as a function of t, (8.7) has distinct roots if and only if t > 5/27 = 0.430... In this case, the black curve has two connected components, and an example is visible in Figure 5 for which θ = π/7. This time the red point (σ, φ) is chosen (with x approximately 4.75) so that there are exactly five

28 points correctly separated from both the (black and red) third and fourth points. Subsequent analysis will show that these five points are not correctly separated from each other. Although the fourth (red) point in Figure 4 is not admissible (nor, in fact, is that in Figure 5), Proposition 6.2 implies that there does exists a SIC set, namely S θ, containing the first three points, so there must be at least six points on the black curve that are admissible. For Figures 6 and 7, we return to the value θ = π/16, and display these six points in green. In Figure 6, we have chosen the fourth point z[σ, φ] to be the admissible one with σ = 0 and φ negative. In Figure 7, we have chosen the fourth point to be one of the points of S θ that does not depend on θ. Recall that the top and bottom boundary of R is a single point, and that the horizontal lines φ = ±π/6 are the circles C 2, C 3. Figure 7 illustrates the fact that any point of C 2 is correctly separated from [z 1 ], [z 2 ] and a given point of C 3, as we explained in the proof of Lemma 8.1. 9. SYMMETRIZATION We suppose now that S is a SIC set containing, in addition to [z 1 ] = [0, 1, ω] and [z 2 ] = [0, 1, ω 2 ], four more points z[σ i, φ i ], i = 3, 4, 5, 6, with t = φ 3, x = φ 4, y = φ 5, z = φ 6. In view of Theorem 7.3 and equation (7.21), our task is to investigate the system of polynomial equations given by (9.1) f(x, y, z) = 0, f(t, y, z) = 0, f(t, x, z) = 0, f(t, x, y) = 0. Since f is itself symmetric, the whole system is invariant under the action of the group of permutations of t, x, y, z. There are refinements of Buchberger s algorithm for dealing with symmetric ideals, but we shall adopt the technique outlined in [34]. Namely, we shall convert the system into a system four equations, each of which involves only the elementary symmetric polynomials defined by (9.2) a = t + x + y + z, b = tx + ty + tz + yz + zx + xy, c = xyz + tyz + txz + txy, d = txyz. To accomplish this, first define (9.3) F 1 = f(x, y, z) + f(t, y, z) + f(t, x, z) + f(t, x, y),